Inhibition of Palm Oil Oxidation by Zeolite Nanocrystals - Journal of

The efficiency of zeolite X nanocrystals (FAU-type framework structure) containing different extra-framework cations (Li+, Na+, K+, and Ca2+) in slowi...
0 downloads 0 Views 5MB Size
Subscriber access provided by University of Birmingham

Article

Inhibition of Palm Oil Oxidation by Zeolite Nanocrystals Kok-Hou Tan, Hussein Awala, Rino R. Mukti, Ka-Lun Wong, Baptiste Rigaud, Tau-Chuan Ling, Hristiyan A. Aleksandrov, Iskra Z. Koleva, Georgi N. Vayssilov, Svetlana Mintova, and Eng-Poh Ng J. Agric. Food Chem., Just Accepted Manuscript • Publication Date (Web): 21 Apr 2015 Downloaded from http://pubs.acs.org on April 22, 2015

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Journal of Agricultural and Food Chemistry is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 37

Journal of Agricultural and Food Chemistry

Inhibition of Palm Oil Oxidation by Zeolite Nanocrystals

1 2 3 4

Kok-Hou Tan,1,# Hussein Awala,2,# Rino R. Mukti,3 Ka-Lun Wong,4 Baptiste Rigaud,2 Tau-

5

Chuan Ling,5 Hristiyan A. Aleksandrov,6 Iskra Z. Koleva,6 Georgi N. Vayssilov,6 Svetlana

6

Mintova,2,* Eng-Poh Ng1,*

7 1

8 2

9

Division of Inorganic and Physical Chemistry, Institut Teknologi Bandung, Indonesia 4

11 12

Laboratoire Catalyse & Spectrochimie, CNRS-ENSICAEN, Université de Caen, France 3

10

5

National Institute of Education, Nanyang Technological University, Singapore

Institute of Biological Sciences, Faculty of Science, University of Malaya, Kuala Lumpur, Malaysia

13 14

School of Chemical Sciences, Universiti Sains Malaysia, Penang, Malaysia

6

Faculty of Chemistry and Pharmacy, University of Sofia, 1126 Sofia, Bulgaria

15 16 17

The efficiency of zeolite X nanocrystals (FAU-type framework structure) containing different

18

extra-framework cations (Li+, Na+, K+ and Ca2+) in slowing down the thermal oxidation of

19

palm oil is reported. The oxidation study of palm oil is conducted in the presence of zeolite

20

nanocrystals (0.5 wt.%) at 180 °C. Several characterization techniques such as visual analysis,

21

colorimetry, rheometry, total acid number (TAN), FT-IR spectroscopy,

22

spectroscopy and Karl-Fischer analyses are applied to follow the oxidative evolution of the

23

oil. It was found that zeolite nanocrystals decelerate the oxidation of palm oil through

24

stabilization of hydroperoxides, which are the primary oxidation product, and concurrently

25

via adsorption of the secondary oxidation products (alcohols, aldehydes, ketones, carboxylic

1

H NMR

1 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 2 of 37

26

acids, and esters). In addition to the experimental results, periodic density functional theory

27

(DFT) calculations are performed to elucidate further on the oxidation process of the palm oil

28

in the presence of zeolite nanocrystals. The DFT calculations show that the metal complexes

29

formed with peroxides are more stable than the complexes with alkenes with the same ions.

30

The peroxides captured in the X zeolite nanocrystals consequently decelerate further

31

oxidation toward formation of acids. Unlike the monovalent alkali metal cations in the X

32

zeolite nanocrystals (K+, Na+ and Li+), the Ca2+ reduced the acidity of the oil by neutralizing

33

the acidic carboxylate compounds to COO–(Ca2+)1/2 species.

34 35

KEYWORDS: Palm oil; oxidation; inhibition; zeolite nanocrystals; extra-framework

36

cations

37 38

INTRODUCTION

39

The overwhelming of palm oil global production (30 % of global vegetable oil output)

40

and trade volume (60 % of global exports) are due to its cheaper production cost and non-

41

toxic nature. The palm oil has been utilized as a base oil in formulating lubricant due to its

42

favorable viscosity-temperature characteristics, high flash point, and compatibility with

43

mineral oil and additive molecules.1,2 However, the low thermal and oxidative stability, poor

44

low-temperature fluidity and hydrolytic instability limit the use of palm oil as long-life

45

cooking oil for deep-frying culinary purposes. Previous studies have been reported that

46

heated cooking oils may pose health risks to consumers due to the generation of oil oxidation

47

products.3-5 Although natural antioxidants can improve the oxidation stability of vegetable oil,

48

but their performances are usually rely very much on the oil composition, heating

49

temperature and the presence of other additives.6 Thus, thermally stable, low cost and

2 ACS Paragon Plus Environment

Page 3 of 37

Journal of Agricultural and Food Chemistry

50

environmentally safe antioxidant additives are desired to replace the traditional antioxidants

51

in inhibiting the oil oxidation particularly in food industry.

52

Zeolites are aluminosilicate microporous solids with well-defined pores and cages,

53

which are commonly used as heterogeneous catalysts, ion-exchangers and adsorbents. In our

54

previous investigation, zeolite nanocrystals (NCs) were used as selective adsorbents for the

55

oil purification.7-9 It is based on the fact that zeolite molecular sieves with electrostatic

56

charged framework tend to adsorb and trap polar compounds (e.g. moisture contaminant and

57

carbonyl oxidized products) in the pores of zeolites without interacting with the non-polar oil

58

molecules. As a result, the oxidized mineral oil after purification can be recycled and re-used

59

with longer service lifetime. Recently, the use of zeolite NCs as eco-friendly anti-oxidant

60

additive was also reported.10,11 Zeolite NCs with different framework type structures (LTL,

61

EMT and FAU) were found to possess distinctive anti-oxidation behavior in palm oil during

62

oxidation at high temperature (180 °C) for a long period of time (40 days).10,11 It was shown

63

that the delay of oxidation progress is the result of free radicals inactivation, C=C bonds

64

stabilization and also adsorption of polar primary (hydroperoxides) and secondary

65

(carboxylic acids) oxidation products, which is greatly depending on the hydrophilicity, pore

66

openings and framework type of the zeolites.

67

Contrary to the well-understood mechanism of chemical reactions contributed by

68

protonated forms of zeolites (Brönsted acids), the chemical properties in cation-exchanged

69

zeolites remains the subject of intense study. The extra-framework cations are compensating

70

the negatively charged framework of zeolites and they are reactive and readily interact with a

71

variety of guest molecules.12,13 This is due to the van der Waals and Coulombic interaction

72

between the extra-framework cations and the guest molecules. The charge, polarizability,

73

deformation ability and influence of non-framework cations on the homogeneity of the

74

framework electric field will eventually affect the sorption and stabilization of diffused

3 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 4 of 37

75

species since these cations are capable to generate strong local electrical fields. Thus

76

attracting the negative centers of polar molecules and polarizing or deforming the polarizable

77

molecules through static electric induction was observed.14,15

78

Synthetic high alumina containing zeolites (FAU, LTA) having alkali and alkaline

79

earth metals as extra-framework cations have been extensively studied, and mainly used as

80

selective adsorbents and for separation of oxygen from air. However, the effect of these

81

cation-exchanged zeolites in oil oxidation has not been investigated. On the other hand, it has

82

been shown that nanosized zeolites prepared via template-free synthesis method do not cause

83

toxicity under cell viability and cell life cycle studies (evaluation of living or dead cells due

84

to toxins was reported).16 Hence these zeolites prepared from template-free precursor

85

suspensions can be considered as additives in palm oil. Furthermore, nanosized zeolites can

86

be incorporated in membranes or deposited as thin film on the culinary tools for deep-frying,

87

apart than been added directly into the cooking oil.

88

Herein, we report the influence of extra-framework cations in nanosized FAU type

89

zeolite (X type) on the oxidation process of palm oil. Nanosized zeolite X samples with four

90

different types of extra-framework cations (Li+, Na+, K+ and Ca2+) are prepared and added as

91

additives during thermal oxidation of the palm oil. The oil oxidative evolution is then

92

characterized and followed by various analytical methods. In addition periodic DFT

93

calculations are performed in order to clarify the effect of extra-framework cations on the

94

chemical interactions and binding energy during the host-guest interactions.

95 96

EXPERIMENTAL SECTION

97

Synthesis and Ion Exchange of Zeolite Nanocrystals. The sodium form of zeolite X

98

(Na-X) nanocrystals (ca. 30 nm) was synthesized without organic template.17 The Li-, K- and

99

Ca-X NCs were prepared through ion exchange method. Typically, 1.50 g of zeolite powder

4 ACS Paragon Plus Environment

Page 5 of 37

Journal of Agricultural and Food Chemistry

100

was stirred in 100 mL of nitrate solution of the targeted metal cations (0.50 mol/L) at 60 °C

101

for 6 h. The process was repeated for 5 times by separating supernatant from mother liquid,

102

re-dispersing in metal nitrate solution and carrying on with aforementioned ion exchange

103

process to ensure that the highest possible ion exchange was achieved. The samples after ion-

104

exchanged were then washed thoroughly (pH = 7.5) prior to freeze-drying.

105

Oxidation of Palm Oil. Zeolite nanocrystals (0.250 g equivalent to 0.5 wt.%) were activated

106

at 180 °C overnight prior adding into 50.00 g palm oil. The oxidation process was carried out

107

at 150 °C under continuous stirring (300 rpm) and reflux; the oil samples were periodically

108

collected after 100 h, 200 h, 300 h and 400 h. The zeolite NCs were recovered from the oils

109

via centrifugation (25000 rpm, 2 h). For comparison, similar amount of palm oil (50.00 g)

110

without zeolite NCs was also oxidized at the same oxidation conditions and denoted as a

111

reference sample (Reference).

112

Characterization. The X-ray diffraction (XRD) patterns of zeolite samples were recorded

113

using a PANalytical X’Pert PRO XRD diffractometer. The morphology of the as synthesized

114

and ion exchanged zeolite nanocrystals was evaluated by field emission scanning electron

115

microscope (FE-SEM) using a Leo Supra 50VP with an accelerating voltage of 30 kV. The

116

porosity of the samples was determined by nitrogen sorption with a Micromeritics ASAP

117

2010 instrument. The chemical composition of all zeolite samples was measured by induced

118

couple plasma using an Agilent 720 Series ICP-OES spectrometer. The IR spectra of zeolites

119

were recorded using a Perkin-Elmer 3000 FTIR spectrophotometer. The thermal stability of

120

the zeolites was measured using a Mettler TGA SDTA851 instrument with a heating rate of

121

10 C/min under nitrogen flow.

122

Evaluation of the Oxidation Degree of Palm Oil. Visual analysis of oil samples was carried

123

out by following their color change after 100 h, 200 h, 300 h and 400 h of oxidation at 150 °C.

124

Colorimetric measurements of oil samples were performed using a Shimadzu UV 3000

5 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 6 of 37

125

spectrophotometer at 530 nm; fresh palm oil was used as a reference. Two measurements

126

were made for each oil sample, and no difference was observed. Qualitative information for

127

the formation of carbonyl compounds, hydroperoxides and moisture in the oil samples was

128

obtained by recording the IR spectra. The IR measurements (50 scans, 4 cm-1) were

129

performed on 0.5 mL of oil introduced in a ZnSe liquid cell (1 mm spacer).

130

The rheological behavior of the oils was investigated by a Malvern Kinexus

131

Rheometer. The analysis was performed twice for each sample and the average values were

132

used for plotting the results. The amount of moisture and organic compounds adsorbed by the

133

zeolite nanocrystals was measured by a Mettler TGA SDTA851 instrument with a heating

134

rate of 10 °C/min under air flow.

135

The water content in oil was measured using a volumetric Karl Fischer titrator. The

136

Karl Fischer titration system (Metrohm) was charged with Hydranal® Composite-2 reactant

137

(Riedel-de Häan), Hydranal® Solvent CM (Riedel-de Häan). All samples were analyzed twice

138

to obtain average values by direct injection without preliminary treatment.

139

Potassium hydroxide (KOH) titrimetric method was applied to determine the Total

140

Acid Number (TAN) of the oil samples. Typically, 0.200 g of oil was mixed with 2.000 g of

141

mixture containing of water : isopropanol : toluene with a molar ratio equal to 1 : 30 : 20,

142

followed by addition of 3 drops of phenolphthalein indicator. The resulting solution was

143

titrated with 0.01 mol/L ethanolic KOH solution until a color change was observed. The

144

results were recorded and the TAN was calculated as mg KOH/g oil using the following

145

equation:

146

TAN (mg KOH/g oil)=

∙∙ 

(Eq. 1)

147

where V is the volume (mL) and M is the molarity (mol/L) of the ethanolic KOH solution, Mr

148

is the molar mass of KOH (g/mol) and m is the mass (g) of the oil sample. Two titrations of

149

each oil sample were performed. 6 ACS Paragon Plus Environment

Page 7 of 37

150

Journal of Agricultural and Food Chemistry

1

H NMR analyses were performed using a Bruker AVIII spectrometer operating at a

151

frequency of 500 MHz with the following acquisition parameters: pulse sequence zg30,

152

acquisition time 3.17 s, relaxation delay 5 s, pulse width 30 for 512 scans. Prior to analysis,

153

the oil samples were dissolved in deuterated chloroform (CDCl3) (mass ratio of zeolite:

154

CDCl3 = 1 : 3); tetramethylsilane (TMS) was used as an internal standard.

155

DFT Modeling of the Interaction of Absorbates with Metal Ions in Zeolite Nanocrystals.

156

Periodic density functional theory (DFT) calculations were performed with the PW91

157

exchange-correlation functional18 using a Vienna ab initio simulation package (VASP).19,20

158

Ultrasoft pseudopotentials were used as implemented in the VASP package.21,22 Due to the

159

large unit cell of the zeolites, the Brillouin zone was sampled using only the Γ point.23 The

160

valence wave functions were expanded in a plane-wave basis with a cutoff energy of 400 eV.

161

The unit cell of the cubic FAU type zeolite was optimized using a pure silica structure

162

with dimensions a = b = c = 24.345 Å.24 During the geometry optimization, all the zeolite

163

atoms and the adsorbate species were allowed to relax until the force on each atom was less

164

than 2×10−4 eV/pm. The binding energy (BE) of the adsorbates hydroperoxides (C2H5OOH

165

and HOOH), and alkenes including ethene (C2H4) and cis-butene (C4H8) per ligand was

166

determined as:

167

BE[adsorbate/Zeo] = {E[adsorbate/Zeo] - E[Zeo] - E[adsorbate]}

168

where E[adsorbate/Zeo] is the energy of the zeolite together with the adsorbed molecule in

169

the optimized geometry; E[Zeo] and E[adsorbate] are the energies of the pure zeolite and

170

pure adsorbate molecule in the gas phase, respectively. With the above definition, the

171

negative values of BE imply favorable interactions.

172 173

RESULTS AND DISCUSSION

7 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 8 of 37

174

Ion Exchanged Zeolite Nanocrystals. The degree of crystallinity and phase purity of FAU

175

zeolite NCs are characterized by XRD. The XRD patterns of all zeolite samples show that the

176

peaks intensity remains intact and no phase transformation occurs, revealing that the degree

177

of crystallinity and phase purity do not change after ion exchange (See Supplementary

178

Information: Figure S1). Besides, the FE-SEM analyses reveal that the size and morphology

179

of zeolite NCs retain stable after ion exchange. All the crystals have a narrow particle size

180

distribution and diameter of 20-40 nm (Figure 1).

181

The schematic structure of zeolite X and the potential positions (I, I’, II, II’, III and

182

III’) to be occupied by its charge compensating cations are shown (ESI: Figure S2) and the

183

results from the chemical analysis of zeolite NCs exchanged with alkali metal and alkali earth

184

metal cations are also listed (Table S1). As can be seen, zeolite X with the FAU framework

185

topology containing pore with a size of 0.74 nm and supercages made up of sodalite cage

186

connected via double 6-rings. The zeolite X nanocrystals have a Si/Al ratio of ca. 1.3. After

187

five cycles of ion exchange at 60 °C, the Li-X, K-X and Ca-X zeolite are successfully

188

prepared while keeping the same Si/Al ratio. In all cases, not 100 % exchange of Na+ is

189

attained (Table S1). The highest degree of exchange of Na-X is achieved for Ca2+ (88 %),

190

followed by K+ (86 %) and Li+ (74 %). Their different ionic radiuses of univalent and

191

divalent cations explain the various degree of ion exchange of Na+ in zeolite X.25 On the

192

other hand, Li-X has the lowest degree of ion exchange, which can be due to its high

193

hydration energy, high mobility, and low occupancy of certain sites.26-28

194

The porosity of zeolite nanocrystals after ion exchange was probed by N2 sorption

195

analysis. All samples exhibits Type I isotherm at low P/P0, which is characteristic for

196

microporous materials (Figure 2). A high adsorption uptake at P/P0 > 0.8 is due to the textural

197

mesoporosity resulting from the close packing of zeolite NCs. From the N2 sorption

198

isotherms, it can be seen that the N2 uptake at low P/P0 is inversely proportional to the size of

8 ACS Paragon Plus Environment

Page 9 of 37

Journal of Agricultural and Food Chemistry

199

cations (Table S1, inset of Figure 2). As a result, Li-X zeolite exhibits the highest BET

200

surface area and pore volumes (both Vmicro and Vmeso), whereas the K-X zeolite containing

201

larger cation K+ (1.33 Å) gives the lowest porosity (Tables S1 and S2). Ca2+ (0.99 Å) is

202

almost similar to Na+ (0.97 Å) in size, but Ca2+ is a divalent ion and the amount of Ca2+ in the

203

Ca-X zeolite is lower. As a result, a slightly increased surface area and pore volume is

204

measured for the Ca-X zeolite. Thus, the results from nitrogen sorption are in agreement with

205

the chemical analysis data, demonstrating that the porosity is slightly affected by the size and

206

the number of cations introduced in zeolite X via ion exchange.

207 208

Characterization of Palm Oil Oxidation

209

Color Change and Colorimetry Analysis of Palm Oil. Color change is a simple indication

210

for oil degradation. Basically, the reference oil containing no zeolites shows faster color

211

change where the oil color progressively turns from yellow to amber, dark brown and finally

212

black color as the severity of the oxidation conditions was increased with time (photo not

213

shown here). The reference oil sample also has a strong smell resulting from volatile organic

214

acids or carbonyl compounds, showing that the oil is degraded. The appearance of oil

215

samples containing zeolite NCs (Na-, Li-, K- and Ca-X) as a function of oxidation time is

216

shown in Figure 3. As can be seen, the oils oxidized with zeolites have brighter color where

217

highly basic K-X displays the brightest color even after 400 h of oxidation. The color change

218

was also found to follow the polarizability/basicity of the cations present in the zeolite X: Li+

219

< Na+ < Ca2+ < K+.

220

The oxidative deterioration of palm oil is always associated with darkening effect.

221

Some organic substances are responsible for the development of color in oil during

222

autoxidation. For example, the oxidation of carotene, which naturally exists in palm oil, can

9 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 10 of 37

223

cause increasing red and yellow colors in vegetable oils.29 Apart from carotene, high

224

molecular weight compounds are also accountable for the color change of palm cooking oil.6

225

The color change was also characterized quantitatively using colorimetry at

226

wavelength, λ = 530 nm where the signal at this wavelength is strongly influenced by the

227

degree of oil deterioration.30

228

As observed by visual analysis, the reference oil evidenced a fast increment in light

229

absorption, and the absorbance value increased tremendously from 3.49 (at 200 h) to 47.11

230

(at 400 h) (Figure 4). The addition of zeolite NCs to the palm oil lower down significantly the

231

level of oil darkness. As shown, the control of absorbance values is highly dependent on the

232

polarizability/basicity of the non-framework cations. Particularly, the oil samples oxidized

233

with highly basic Ca-X and K-X zeolite NCs exhibit much slower increment in absorbance

234

intensity before 300 h, where only about 0.95 of light absorbance was recorded. After 400 h

235

of oxidation, the palm oil with Ca-X zeolite has the lowest absorbance value of 4.89, whereas

236

the palm oil containing K-X has an absorbance value of 5.20. On the other hand, less basic

237

Li-X and Na-X were less effective in controlling the oil darkening; an absorbance value of

238

7.15 and 5.64 was measured, respectively, after 400 h of oxidation.

239

Viscosity and Acidity Changes of Palm Oil. Rheometry was used to determine the fluidity

240

of the palm oil samples as a function of oxidation time (Figure 5). At the beginning, the fresh

241

oil had a viscosity value as low as 72 cSt. With increasing the oxidation time, each oil sample

242

experienced different extent of viscosity elevation. The rate of oil thickening was

243

significantly reduced when zeolite X NCs were added. For instance, the viscosity of palm oil

244

with K-X NCs increased slowly to 912 cSt after 400 h of oxidation, which was 5 times lower

245

than that of the oxidized reference oil (4426 cSt). The rate of viscosity change with time was

246

also calculated based on the slope of the original function (dη/dt), where this derivative

247

enables to correlate cation properties with oil viscosity at any instant (Figure 5b). The results

10 ACS Paragon Plus Environment

Page 11 of 37

Journal of Agricultural and Food Chemistry

248

show that the control of oil viscosity in the presence of zeolite NCs is depending on the

249

cations charge density and polarizability. For example, larger extra-framework cations with

250

low charge density such as K+ (0.101 e/Å3) and Na+ (0.268 e/Å3) have better performance in

251

resisting to the change in viscosity of palm oil than Li+ (0.779 e/Å3) and Ca2+ (0.505 e/Å3).31

252

Thus, the results show that the oil oxidized with K-X zeolite has the lowest viscosity due to

253

the presence of low concentration of polymeric oxidized compounds.

254

Furthermore, the oil viscosity and oil darkening are also related with the oil acidity

255

because the presence of polar acidic groups (oxidized products) tends to enhance the polarity

256

of oil molecules, thus attracting the organic compounds and resulting in bulkier molecules.

257

Typically, the acidic compounds are mainly alcohols, aldehydes, lactones and carboxylic

258

acids. These oxidation products, which contain acidic protons, are formed as a result of

259

scission reactions of unsaturated fatty acid chain during oxidative chain reaction.32 The

260

acidity of oil was evaluated based on the Total Acid Number (TAN) measurements of the oil

261

samples. As expected, the trend of oil acidity is almost similar to the trend of viscosity of oil,

262

where the reference oil shows fast increment in the TAN value, i.e. 46.83 mg KOH/g oil of

263

TAN was recorded after 400 h (Figure 6a). In contrast, the oils with zeolite NCs exhibit slow

264

rise in acidity. The acidity of the oils containing zeolite nanocrystals is approximately 4 times

265

lower than that of reference oil. The derivative measurement of the instantaneous changes in

266

the TAN value of oil samples was also performed, and the plot is shown in Figure 6b. It can

267

be clearly seen that the oils oxidized with highly basic K-X zeolite has the lowest rate of

268

TAN values change, indicating that K-X is the most effective in controlling the oil

269

deterioration. While, the basic divalent cation in sample Ca-X has comparable performance to

270

sample Na-X in halting oil oxidation, followed by Li-X. Thus, the similarity of trend in the

271

TAN plot to those of colorimetry and rheometry provide a strong proof to the existence of

11 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 12 of 37

272

relationship between the oil composition, the rate of oil darkening and the total acidic

273

compounds produced, which is consistent with our previous finding.10

274

Oil Oxidation Followed by FTIR and 1H NMR Spectroscopy. FTIR spectroscopy was

275

further applied to study the development of carbonyl-oxidized products in the oil (vibrations

276

in the range 1900‒1400 cm-1). Mainly, four signals are slowly developed at 1784 cm-1, 1743

277

cm-1, 1703 cm-1 and 1655 cm-1 during the oxidation process of oil, which correspond to

278

lactones, esters, carboxylic acids and water, respectively (inset of Figure 7). The reference

279

sample (oil without zeolite) experiences fast peak broadening and significant baseline offset

280

at 1680‒1740 cm-1 during the oxidation process (inset of Figure 7). Integration of the peak

281

area (1690‒1725 cm-1) is used to provide a comprehensive view to the quantitative evolution

282

of carboxylic acids along the oxidation for 400 h (Figure 7). The result reveals a sharp

283

increase in peak area in the reference oil especially after 200 h of oxidation indicating the fast

284

development of carbonyl-oxidized compounds in the palm oil. The addition of zeolite NCs,

285

slows down oxidation, and the broadening of the IR peak at 1700-1800 cm-1, is indicating

286

deceleration of carboxylic acids in the oil. The effect of extra-framework cations in halting

287

oil oxidation is followed; the changes in the integrated peak area at 1700-1720 cm-1 show the

288

dependency of carboxylic acid generation on the types of extra-framework cations in the

289

zeolite nanocrystals. As can be seen, the production rate of carboxylic acid is directly

290

proportional to the polarizability of the alkali metal cations (Figure 7). This is evidenced by

291

the steeper slope of the curves obtained from the integrated peak area for oil samples treated

292

with Li-X than those with K-X and Ca-X. Among all oil samples, the one containing K-X has

293

the least steep slope suggesting that the oxidative inhibition of the zeolite improves with

294

decreasing basicity.

295

Additionally, the formation of hydroxyl compounds in the oil under oxidation is

296

estimated using the FTIR data. The absorption band of alcohols (ca. 3615 cm-1), water (ca.

12 ACS Paragon Plus Environment

Page 13 of 37

Journal of Agricultural and Food Chemistry

297

3534 cm-1) and hydroperoxides (ca. 3468 cm-1) are overlapping, thus only indication for the

298

overall development of these compounds in the oil samples is observed (ESI: Figure S3). The

299

IR results reveal that the zeolites NCs are controlling the rate of formation of hydroxyl

300

compounds in comparison to the oil free of zeolite (reference) during the oxidation process.

301

The formation rate of hydroxyl compounds in the palm oil with Li-X is higher than those of

302

Na-X, K-X and Ca-X oil samples, indicating that Ca-X and K-X are able to hinder the

303

formation of hydroxyl compounds.

304

Water is an undesirable oxidized byproduct because it catalyzes oil oxidation and

305

promotes acid formation. Hence, the water content in the oil samples is evaluated using the

306

area of the IR peaks in the region 3500-3350 cm-1. As can be seen, the peaks areas in this

307

region increase with oxidation time, which is a clear indication for increase of water content

308

during the oxidation process (ESI: Figure S3). Nevertheless, this region is not sensitive and

309

accurate to determine the water content in oil media due to the disrupted O-H (alcohols,

310

carboxylic acids and hydroperoxides) stretching vibration modes.33 Therefore, the moisture

311

content was quantitatively measured with Karl Fischer titration method.

312

The initial water content in the fresh palm oil is 826 ppm, which is less than 0.1 % of

313

the mass of cooking oil – a desired level of water content according to Palm Oil Refiners

314

Association of Malaysia (PORAM) specification (Figure 8).34 As can be seen, in the

315

reference oil the water content during the oxidation process is increasing fast. This can be

316

explained by the high oxidation rate of the oil where water is produced as a degradation by-

317

product via condensation. In contrary, the oil samples containing zeolite NCs overall contain

318

lower water content than the reference oil during the oxidation period. The moisture content

319

in palm cooking oil is dropped during the first 200 h, which might be due to the adsorption of

320

water in the zeolites. Since then, the quantity of water slowly increases, depending on the

321

type of counter ions of zeolite X. The lowest water content after 400 h is detected in the oil

13 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 14 of 37

322

with Ca-X (1498 ppm), while the oil with Li-X has the highest level of moisture (1696 ppm).

323

These changes in water content can be explained by the effect of controlled thermal oxidation

324

process. The better the zeolite in reducing the rate of oxidation, the slower the generation of

325

water, and hence the lower amount of water is present in the oil.

326

Furthermore, a detail study on the formation of carbonyl compounds including

327

carboxylic acid (>9.6 ppm) and aldehydes (9.3 ppm‒9.6 ppm) using 1H NMR spectroscopy is

328

performed. It is demonstrated that the cations with lower polarizability/lower basicity are

329

acting less effectively in slowing down the formation of the acidic components in the palm

330

oil (Figure 9). For instance, the formation of aldehydes and carboxylic acids in the palm oil

331

oxidized in the presence of Li-X is faster than the oil containing more basic cations. Among

332

all samples, the palm oil containing K+ with high cation polarizability has the least increment

333

in both aldehyde and carboxylic contents, followed by the divalent Ca-X in halting oil

334

oxidation possess.

335

The different colors in oil samples can be due to the presence of high-molecular-

336

weight polymeric oxidized compounds via intensive light absorption and scattering effect. In

337

addition, different content of beta-carotene will also lead to different colors in oxidized oil

338

samples as well. Hence, 1H NMR spectroscopy was used to detect the presence of beta-

339

carotene in oil samples after 300 h of oxidation (ESI: Figure S4). These oil samples were

340

selected since they exhibited colors significantly different from their counterparts. The 1H

341

NMR spectra show the characteristics peaks of beta-carotene at 6.77 ppm arising from

342

protons 11/11' (6.76 ppm) and protons 15/15' (6.74 ppm). In addition, protons 7/7' (6.24 ppm)

343

together with protons 10/10' (6.23 ppm) show an overlapping signal at 6.24 ppm. The protons

344

8/8' is found at 6.19 ppm.35 It can be seen that the peaks intensity of carotene in the oils

345

oxidized with zeolite NCs is almost identical to that of reference oil, indicating that the

14 ACS Paragon Plus Environment

Page 15 of 37

Journal of Agricultural and Food Chemistry

346

oxidation rate of carotene is not affected by the presence of zeolite NCs. Thus, taking this fact

347

into consideration, the change of oil color is mainly due to the oxidation products.

348

Besides, the content of hydroperoxides in the palm oils is also followed by 1H NMR

349

spectroscopy; the signal at ca. 8.80 ppm corresponds to hydroperoxide group (RO–OH).36 It

350

can be seen that no peak is observed in the reference oil throughout the oxidation process

351

showing that the hydroperoxides (primary oxidized products) is very unstable and they are

352

directly converted into carboxylic acids, aldehydes, etc. (secondary oxidized products) (ESI:

353

Figure S5). As a result, the oxidation rate and acidity of the palm oil increases. In contrast, a

354

broad peak at 8.8 ppm is observed in all oils samples containing zeolite NCs during the first

355

200 h, indicating their stabilization in the zeolite NCs (ESI: Figure S6). Thus, less secondary

356

oxidized products (e.g. carboxylic acids and aldehydes) and lower TAN values are observed

357

in these samples. Among four basic exchanged cations, the Ca2+ with the highest

358

polarizability contributes the most significantly to the stabilizing of hydroperoxides, followed

359

by K+, Na+ and Li+. The broad peak at 8.8 ppm, however, completely disappears after 300 h,

360

and the oxidation rate increases as shown by colorimetry, TAN, rheometry and IR

361

spectroscopic studies. This can be explained by competitive adsorption of primary and

362

secondary oxidation products. The mechanism of oxidative inhibition of palm oil by zeolite

363

NCs will be discussed further in Section 3.5.

364

Characterization of Zeolite Nanocrystals after Oil Oxidation. The zeolite nanocrystals

365

used during oil oxidation were separated from the oil and washed before subjected to

366

characterization. Figure S7 shows the IR spectra of zeolite samples before and after used in

367

oil oxidation. It can be seen that fresh Na-X has several peaks at 1006, 752, 670, 566, 459

368

cm-1 characteristic for the framework structure of zeolite X, and the peaks at 3410 and 1647

369

cm-1 are due to adsorbed water. After oil oxidation, the IR bands of zeolite NCs remain

370

unchanged indicating that the framework structure of all zeolite samples does not change

15 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 16 of 37

371

after 400 h of oxidation. In addition, several peaks are observed at 2921, 2857, 1462 and

372

1363 cm-1, which are attributed to the adsorption of organic molecules by the zeolite NCs.

373

Furthermore, the hydroxyl band at 3410 cm-1 is broadened and slightly shifted to 3405 cm-1

374

revealing that the zeolites adsorb also hydrocarbons containing hydroxyl groups such as

375

alcohols and hydroperoxides. For zeolite NCs with monovalent cations, a signal at 1745 cm-1

376

that corresponds to esters is overlapping with the signal of 1650 cm-1 (OH bending). Thus,

377

the IR results conclude that the oxidative inhibition in oil by monovalent cation exchanged

378

zeolite is not due to oxidation products removal action because no carboxylic acids are

379

adsorbed (no peak is found at 1715 cm-1). On the other hand, the zeolite Ca-X containing

380

divalent cations exhibits a strong peak at 1599 cm-1, which is corresponding to the COO-

381

species.

382

All zeolite used during the oxidation of palm oil are studied by TG/dTG analysis.

383

Basically, all samples display two steps of weight losses. The zeolites adsorb almost similar

384

amount of water (5.0 %), which is desorbed at 160 C (ESI: Figure S8a). However, the

385

amount of organic compounds adsorbed and their desorption temperature vary depending on

386

the type of zeolite nanocrystals. The Li-X with the largest surface area and pore volume

387

adsorbs the highest amount of organic compounds (35.0 %) followed by Na-X (28.0 %) and

388

K-X (23.0 %). The temperature of desorption is identifying the species adsorbed and also

389

provides information on the strength of interaction between that compound and the counter

390

cations of zeolite. The dTG curve of Li-X exhibits two signals at 275 °C and 338 °C (ESI:

391

Figure S8b) whereas only one signal around 301‒310 °C for samples Na-X and K-X is

392

observed (ESI: Figure S8b). As seen, the desorption temperature for K-X is slightly lower

393

than that of Na-X indicating that K-X has weaker interaction with the organic compounds due

394

to its lower charge density. On the other hand, divalent cations in sample Ca-X adsorb about

395

30 % of organic compounds, which is almost similar to Na-X. Three signals at 275 °C, 387

16 ACS Paragon Plus Environment

Page 17 of 37

Journal of Agricultural and Food Chemistry

396

°C and 433 °C for sample Ca-X are observed, showing that the Ca2+ has higher affinity to

397

organic species than monovalent cations.

398

Modeling of the Interaction of Peroxide and Alkene Complexes with Metal Ions in

399

Zeolite Nanocrystals. Counter cations in the zeolites can interact with C=C bonds and

400

hydroperoxides during the oil oxidation process. Therefore we consider adsorption of

401

simplified hydroperoxides (e.g. HOOH and C2H5OOH) and alkenes (e.g. ethene and cis-

402

butene) on the M-X zeolites (M = Li, Na, K, Ca). The calculations show that the counter

403

cations introduced in the zeolite X more preferably interact with the hydroperoxides than

404

with the alkenes. As shown in Table S3, the binding energy of peroxides (from –50 to –109

405

kJ/mol) with the metal cations is much higher than that of the binding energy of alkenes

406

(from –17 to –69 kJ/mol), which clearly indicates that the extra-framework cations of zeolite

407

X tend to stabilize peroxides. This result is in good agreement with the 1H NMR spectroscopy

408

data where large concentration of hydroperoxides is detected in the oil containing zeolite

409

nanocrystals during the first 200 h of oxidation and thus, the decomposition rate of

410

hydroperoxide is decelerated due to the stabilization of hydroperoxides by the extra-

411

framework cations (ESI: Figure S6). On the other hand, basic counter cations have lower

412

tendency to interact with the alkenes as indicated by their lower binding energy (from –17 to

413

–69 kJ/mol). An attempt to investigate the effect of addition of methyl and ethyl groups in the

414

alkene molecules on the binding energy is made. The DFT calculations reveal that the

415

insertion of alkyl groups in the alkene chain has insignificant effect (C=C< + ROO•

>C=O + >C=O + R•

(5)

457

Polymerization

: R• + R’• → R‒R’ (high molecular weight compounds)

(6)

458

and Termination

: ROO• + R• → ROOR

(7)

459 460

When the zeolite NCs containing different alkali metal cations are added during oil

461

oxidation, the oxidation process pathway is altered, and the oxidation rate is slowed down.

462

Initially, the hydrocarbons are oxidized to hydroperoxides. In the presence of zeolites, the

463

hydroperoxides are stabilized by the extra-framework cations and thus, the decomposition

464

rate of the hydroperoxides is decelerated (Equations 2a and 3a). This process is confirmed by

465

the 1H NMR spectroscopy results and supported by the periodic DFT calculations. The palm

466

oil with zeolite NCs have large amount of hydroperoxides as proven by the presence of a

467

broad signal at 8.8 ppm in the 1H NMR spectra. As a result, this reaction becomes a limitation

468

step to inhibit the oxidation reaction. Thus, a slow oxidation process of the palm oil is

469

observed during the first 200 h by colorimetry, rheometry, TAN and IR spectroscopy

19 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 20 of 37

470

analyses (Figures 4‒7). Among the exchanged cations, the divalent Ca2+ with the highest ion

471

polarizability stabilizes the peroxides better than the monovalent K+, Na+ and Li+ cations. The stabilizing effect by extra-framework cations, however, becomes weaker and the

472 473

1

474

oxidation and polymerization increase after 300 h. At this stage, the zeolites are also

475

behaving as oxidative inhibitors by adsorbing oxidation products such as alcohols,

476

hydroperoxides, esters, water, etc. Unlike monovalent alkali metal cations exchanged zeolites,

477

the alkaline earth Ca-X also reduced the acidity of the oil by neutralizing the acidic

478

carboxylate compounds to COO–(Ca2+)1/2 species as proven by IR spectroscopy (ESI: Figure

479

S7) and DFT simulations (Figure 10, Table S3).

H NMR signal at 8.8 ppm completely disappears after 300 h (ESI: Figure S6). Hence, the oil

480 481

CONCLUSIONS

482

This work reveals the effect of extra-framework cations in zeolite X nanocrystals on

483

the oxidation of palm oil. The results show that the anti-oxidation behaviour of zeolite

484

nanocrystals depends on basicity, charge density and polarizability of the counter cations of

485

the zeolites. Li-X zeolite nanocrystals with the lowest cation polarizability are found to have

486

the lowest oxidative inhibition activity, while the highly basic and highly polarized K-X

487

zeolite nanocrystals are the best candidate to hinder oil oxidation. The effect of zeolite

488

nanocrystals on oil oxidation is found to be different for mono- and divalent cations

489

containing materials. According to 1H NMR spectroscopy and periodic DFT calculation, all

490

counter ions, especially Ca2+ with bidentate property, are able to stabilize the hydroperoxides,

491

which are the primary oxidation products, and hence, the decomposition rate of

492

hydroperoxides and the propagation step are decelerated. Furthermore, the adsorption is also

493

another mechanism in halting oil degradation by removing the harmful compounds such as

494

esters, alcohols, water and hydroperoxides as revealed by IR spectroscopy. For Ca-X zeolite

20 ACS Paragon Plus Environment

Page 21 of 37

Journal of Agricultural and Food Chemistry

495

nanocrystals, the bidentate capability of Ca2+ leads to neutralization of the acidic carboxylate

496

compounds that reduces the acidity of the oil and further decelerates of oil oxidation process.

497 498 499

ASSOCIATED CONTENT

500

Supporting Information

501

X-ray diffraction patterns of zeolite samples; FTIR spectra of oil samples; 1H NMR spectra of

502

palm oil (6.0-7.2 ppm and 8.5-9.3 ppm) oxidized at 150 °C; 1H NMR spectra of the palm oil

503

oxidized in the presence of zeolites; FTIR spectra of zeolite samples; TGA and dTG of

504

zeolite nanocrystals after 400 h of oxidation. These materials are available free of charge via

505

the Internet at http://pubs.acs.org.

506

Corresponding Authors

507

*

Email (SM): [email protected]; (EPN): [email protected]

508

#

These authors contributed equally to this work as first author.

509 510

ACKNOWLEDGEMENTS

511

The authors would like to acknowledge Bio-Asia Program and FRGS (203/PKIMIA/6711362)

512

research grants for financial support. Kok-Hou Tan would also like to thank the MyBrain and

513

USM fellowship for the scholarship provided.

514 515

REFERENCES

516

(1) Carter, C.; Finley, W.; Fry, J.; Jackson, D.; Willis, L. Palm Oil Markets and Future

517 518 519

Supply. Eur. J. Lipid Sci. Technol. 2007, 109, 307-314. (2) Sharma, B. K.; Adhvaryu, A.; Liu, Z.; Erhan, S. Z. Chemical Modification of Vegetable Oils for Lubricant Applications. J. Am. Oil Chem. Soc. 2006, 83, 129-136.

21 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 22 of 37

520

(3) Takeoka, G. R.; Full, G. H.; Dao, L. T. Effect of Heating on the Characteristics and

521

Chemical Composition of Selected Frying Oils and Fats. J. Agric. Food Chem. 1997,

522

45, 3244-3249.

523

(4) Billek, G. Health Aspects of Thermoxidized Oils and Fats. Eur. J. Lipid Sci. Technol. 2000, 102, 587-593.

524 525

(5) Ebong, P. E.; Owu, D. U.; Isong, E. U. Influence of Palm Oil (Elaesis guineensis) on Health. Plant Foods Hum. Nutr. 1999, 53, 209-222.

526 527

(6)

Oxidation. Tribol. Int. 2007, 40, 1035-1046.

528 529

(7)

(8)

Majano, G.; Ng, E. -P.; Mintova, S. Inhibition of Lubricant Degradation by Nanoporous Materials. Stud. Surf. Sci. Catal. 2008, 174, 569-572.

532 533

Ng, E.-P.; Delmotte, L.; Mintova, S. Selective Capture of Water using Microporous Adsorbents to Increase the Lifetime of Lubricants. ChemSusChem 2009, 2, 255-260.

530 531

Fox, N. J.; Stachowiak, G. W. Vegetable Oil-Based Lubricants - A Review of

(9)

Ng, E. -P.; Mintova, S. Nanoporous Materials as High Selective Water Sorbents in

534

Lubricants. In Topics in Chemistry and Material Science: Advanced Micro- and

535

Mesoporous Materials; Hadjiivanov, K., Valtchev, V., Mintova, S., Vayssilov, G.,

536

Eds.; Heron Press Science Series: Bulgaria, 2008; Vol. 1, pp 322-329.

537

(10) Tan, K. -H.; Ng, E. -P.; Awala, H.; Mukti, R. R.; Wong, K. -L.; Yeong, S. -K.; Mintova,

538

S. Halting Oxidation of Palm Oil with Eco-friendly Zeolites. In Topics in Chemistry

539

and Material Science: Advanced Micro- and Mesoporous Materials; Hadjiivanov, K.,

540

Valtchev, V., Mintova, S., Vayssilov, G., Eds.; Heron Press Science Series: Bulgaria,

541

2014; Vol. 7, pp 102-108.

542

(11) Majano, G.; Ng, E.-P.; Lakiss, L.; Mintova, S. Nanosized Molecular Sieves Utilized as

543

An Environmentally Friendly Alternative to Antioxidants for Lubricant Oils. Green

544

Chem. 2011, 13, 2435-2440.

22 ACS Paragon Plus Environment

Page 23 of 37

Journal of Agricultural and Food Chemistry

545

(12) Mikosch, H.; Uzunova, E. L.; Nikolov, G. S. Interaction of Molecular Nitrogen and

546

Oxygen with Extraframework Cations in Zeolites with Double Six-membered Rings

547

of Oxygen-bridged Silicon and Aluminum Atoms: A DFT Study. J. Phys. Chem. B

548

2005, 109, 11119-11125.

549

(13) Sethia, G.; Somani, R. S.; Bajaj, H. C. Adsorption of Carbon Monoxide, Methane and

550

Nitrogen on Alkaline Earth Metal Ion Exchanged Zeolite-X: Structure, Cation

551

Position and Adsorption Relationship. RSC Adv. 2015, 5, 12773-12781.

552

(14) Cozens, F. L.; Cano, M. L.; García, H.; Schepp, N. P. Alkali Metal Cation Control of

553

Oxidation Reactions of Radicals in Zeolites. J. Am. Chem. Soc. 1998, 120, 5667-5673.

554

(15) Xu, R.; Pang, W.; Yu, J.; Huo, Q.; Chen, J. Chemistry of zeolites and related porous

555

materials: Synthesis and Structure; John Wiley & Sons (Asia) Pte Ltd: Singapore,

556

2007; pp 354.

557

(16) Laurent, S.; Ng, E.-P.; Thirifays, C.; Lakiss, L.; Goupil, G.-M.; Mintova, S.; Burtea, C.;

558

Oveisi, E.; Hébert, C.; de Vries, M.; Motazacker, M. M.; Rezaee, F.; Mahmoudi, M.

559

Corona Protein Composition and Cytotoxicity Evaluation of Ultra-small Zeolites

560

Synthesized from Template Free Precursor Suspensions. Toxicol. Res. 2013, 2, 270-

561

279.

562

(17) Awala, H.; Gilson, J-P.; Retoux, R.; Boullay, P.; Goupil, J-M.; Valtchev V.; Mintova S.

563

Template-free Nanosized Faujasite Type Zeolites. Nature Mater. 2015, doi:

564

10.1038/nmat4173.

565 566

(18) Perdew, J. P.; Wang, Y. Accurate and Simple Analytic Representation of the ElectronGas Correlation Energy. Phys. Rev. B 1992, 45, 13244-13249.

567

(19) Kresse, G.; Hafner, J. Ab Initio Molecular-dynamics Simulation of the Liquid-Metal

568

Amorphous-Semiconductor Transition in Germanium. Phys. Rev. B 1994, 49,

569

14251-14269.

23 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 24 of 37

570

(20) Kresse, G.; Furthmüller, J. Efficiency of Ab-initio Total Energy Calculations for Metals

571

and Semiconductors using a Plane-Wave Basis Set. Comput. Mater. Sci. 1996, 6, 15-

572

50.

573 574 575 576

(21) Vanderbilt, D. Soft Self-consistent Pseudopotentials in a Generalized Eigenvalue Formalism. Phys. Rev. B 1990, 41, 7892-7895. (22) Kresse, G.; Hafner, J. Norm-conserving and Ultrasoft Pseudopotentials for First-row and Transition Elements. J. Phys.: Condens. Matter 1994, 6, 8245–8257.

577

(23) Jeanvoine, Y.; Ángyán, J. G.; Kresse, G.; Hafner, J. Brønsted Acid Sites in HSAPO-34

578

and Chabazite: An Ab Initio Structural Study. J. Phys. Chem. B 1998, 102, 5573–

579

5580.

580 581

(24) Database of Zeolite Structures. http://www.iza-structure.org/databases/ (accessed Nov 8, 2014).

582

(25) Yang, Y.; Burke, N.; Zhang, J.; Huang, S.; Lim, S.; Zhu, Y. Influence of Charge

583

Compensating Cations on Propane Adsorption in X Zeolites: Experimental

584

Measurement and Mathematical Modeling. RSC Adv. 2014, 4, 7279-7287.

585 586

(26) Sherry, H. S. Handbook of Zeolite Science and Technology; Marcel Dekker: New York, 2003; pp 1007-1061.

587

(27) Feuerstein, M.; Accardi, R. J.; Lobo, R. F. Adsorption of Nitrogen and Oxygen in the

588

Zeolites Li-A and Li-X Investigated by 6Li and 7Li MAS NMR Spectroscopy. J.

589

Phys. Chem. B 2000, 104, 10282-10287.

590

(28) Zhu, J.; Mosey, N.; Woo, T.; Huang, Y. Study of the Adsorption of Toluene in Zeolite

591

LiNa−Y by Solid-State NMR Spectroscopy. J. Phys. Chem. C 2007, 111, 13427-

592

13436.

593 594

(29) Golumbic, C. Antioxidants and Autoxidation of Fats. XIV. The Isolation of New Antioxidants. J. Am. Chem. Soc. 1942, 64, 2337-2340.

24 ACS Paragon Plus Environment

Page 25 of 37

595 596 597 598

Journal of Agricultural and Food Chemistry

(30) Engelsen, S. B. Explorative Spectrometric Evaluations of Frying Oil Deterioration. J. Am. Oil Chem. Soc. 1997, 74, 1495-1508. (31) CRC Handbook of Chemistry and Physics. http://www.hbcpnetbase.com (accessed Nov 8, 2014).

599

(32) Schaich, K. M. Lipid Oxidation: Theoretical Aspects. In Bailey's Industrial Oil and Fat

600

Products; Shahidi, F., Eds.; John Wiley & Sons, Inc.: New Jersey, 2005; Vol. 1, pp

601

269-357.

602

(33) Ng, E.-P.; Mintova, S. Quantitative Determination of Moisture in Lubricating Oils by

603

FT-IR Spectroscopy Combined with Solvent Extraction Approach. Microchem. J.

604

2011, 98, 177-185.

605

(34) PORAM

Standard

Specifications

for

Processed

Palm

Oil,

606

http://poram.org.my/contracts/poram-standard-specifications-for-processed-palm-

607

oil/ (accessed Nov 8, 2014).

608

(35) Holtin, K.; Albert, K. Chapter 4: The Use of NMR Detection of LC in Carotenoid

609

Analysis. In Carotenoids: Physical, Chemical, and Biological Functions and

610

Properties; Landrum, J. T.; Ed.; CRC Press: Florida, 2009; pp. 61-74.

611

(36) Turovskij, N. A.; Berestneva, Y. V.; Raksha, E. V.; Zubritskij, M. Y.; Grebenyuk, S. A.

612

NMR Study of the Complex Formation between Tert-butyl Hydroperoxide and

613

Tetraalkylammonium Bromides. Monatsh. Chem. 2014, 145, 1443-1448.

614 615 616 617 618 619

25 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

620

Page 26 of 37

Figures

621

622 623

Figure 1. SEM images of (a) Li-X, (b) Na-X, (c) K-X, and (d) Ca-X zeolite nanocrystals.

624

Scale bar M = 500 nm.

625 626

26 ACS Paragon Plus Environment

Page 27 of 37

Journal of Agricultural and Food Chemistry

627 628

Figure 2. Nitrogen adsorption (close symbols) and desorption (open symbols) isotherms of

629

zeolites Li-X, Na-X, K-X and Ca-X.

630 631

27 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 28 of 37

632 633

Figure 3. Color change of palm oil with increasing oxidation time (a) 100 h, (b) 200 h, (c)

634

300 h and (d) 400 h.

635 636 637 638

28 ACS Paragon Plus Environment

Page 29 of 37

Journal of Agricultural and Food Chemistry

639 640

Figure 4. Colorimetric evaluation of palm oil samples oxidized with and without zeolites Li-

641

X, K-X, Na-X, and Ca-X.

642 643 644

29 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 30 of 37

645 646

Figure 5. (a) Viscosity of palm oil samples with or without zeolites Li-X, K-X, Na-X, and

647

Ca-X with oxidation time, and (b) plot of dη/dt versus time showing the increasing rate of

648

viscosity of the oil samples.

649 650

30 ACS Paragon Plus Environment

Page 31 of 37

Journal of Agricultural and Food Chemistry

651 652

Figure 6. (a) Development of TAN and (b) differential function of TAN of palm oil samples

653

with or without zeolites Li-X, K-X, Na-X, and Ca-X with oxidation time. Inset in (b)

654

magnification of the plot is shown.

655 656

31 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 32 of 37

657 658

Figure 7. Development of integrated IR peak area (1690-1725 cm-1) corresponding to the

659

carboxylic acids formed during the oxidation of palm oil with and without zeolites Li-X, K-X,

660

Na-X, and Ca-X. Inset: FTIR spectra of reference oil in the range of 1550-1900 cm-1. The

661

pink area indicates the region used for integration of peak area.

662

32 ACS Paragon Plus Environment

Page 33 of 37

Journal of Agricultural and Food Chemistry

663 664

Figure 8. Change in water content of oxidized palm oil samples with and without zeolites Li-

665

X, K-X, Na-X, and Ca-X with oxidation time.

666 667 668

33 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 34 of 37

669 670

Figure 9. 1H NMR spectra of palm oil samples oxidized with and without zeolites Li-X, K-X,

671

Na-X, and Ca-X for (a) 100 h, (b) 200 h, (c) 300 h and (d) 400 h.

672 673 674 675 676

34 ACS Paragon Plus Environment

Page 35 of 37

Journal of Agricultural and Food Chemistry

677 678

Figure 10. Optimized structures in periodic models. Local structures of selected complexes:

679

(a) Ca2+(CH3CH2OOH) and (b) Na+(CH3CH2OOH). (c) Location of the complexes in the

680

cavity of the FAU type structure, represented by Ca2+(CH3CH2OOH). Color coding: red – O,

681

gray – Si, blue – Al, white – H, green – Ca, light blue – Na.

682 683 684 685 686 687 688 689

35 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 36 of 37

690 691 692

Table of Contents (TOC) graphic

693

36 ACS Paragon Plus Environment

Page 37 of 37

Journal of Agricultural and Food Chemistry

 

ACS Paragon Plus Environment